Microstructure dependence of diffusional transport

Microstructure dependence of diffusional transport

Computational Materials Science 20 (2001) 37±47 www.elsevier.com/locate/commatsci Microstructure dependence of di€usional transport Jingzhi Zhu a,*,...

970KB Sizes 2 Downloads 79 Views

Computational Materials Science 20 (2001) 37±47

www.elsevier.com/locate/commatsci

Microstructure dependence of di€usional transport Jingzhi Zhu a,*, Long-Qing Chen a, Jie Shen b, Veena Tikare c a

Department of Materials Science and Engineering, 231 Steidle Building, Penn State University, University Park, PA 16802, USA b Department of Mathematics, Penn State University, University Park, PA 16802, USA c Materials Modeling and Simulation, Sandia National Laboratory, Albuquerque, NM 87185-1411, USA Received 2 February 2000; received in revised form 10 May 2000; accepted 30 May 2000

Abstract A simple and e€ective numerical method is proposed for simulating the temporal di€usive mass transport process through a microstructure with arbitrary complexity described by a phase-®eld approach. The mass di€usion through a given microstructure is modeled by a di€usion equation with a variable di€usion coecient, which is solved by an ecient and accurate semi-implicit spectral method. It is shown that it is possible to extract the e€ective di€usion coecient for any given microstructure from the temporal concentration pro®les. The method is used to simulate the grain boundary di€usion in a single-phase polycrystalline grain structure and the heterogeneous di€usion in a twophase microstructure with di€erent di€usion coecient in each phase. Results are compared with existing analytical theories and computer simulations. Ó 2001 Elsevier Science B.V. All rights reserved. PACS: 66.30.-h; 02.60.Cb; 66.10.Cb Keywords: Di€usion; Microstructures; Phase-®eld; Spectral method; Grain boundary di€usion; E€ective di€usion coecient

1. Introduction Di€usion in a solid with a multi-phase microstructure or with defects such as grain boundaries and dislocations can be highly heterogeneous. For example, in a polycrystalline grain structure, the di€usion coecients along a grain boundary and through a bulk lattice can be orders of magnitude di€erent. As a result, although the volume fraction of grain boundaries is small, grain boundary diffusion may dominate a number of processes such as grain growth, micro-electronics device failure,

* Corresponding author. Tel.: +1-814-863-9957; fax: +1-814867-0476. E-mail address: [email protected] (J. Zhu).

and sintering [1]. Macroscopically, the e€ective di€usion coecient of a grain structure depends on the relative magnitudes of the grain boundary and bulk di€usion coecients, the average grain size and grain boundary width. Although it is possible to derive analytical solutions for very special microstructures, it is impossible to derive the e€ective properties of a microstructure exactly. Therefore, most of the statistical theories are concerned with the property bounds [2]. Recently, there has been increasing interest in computing the e€ective properties from a microstructure either from a digitization of an experimental microstructure or from a microstructure simulation model [3]. However, most of these approaches are based on steady-state equations and thus cannot describe the temporal mass di€usive transport behavior.

0927-0256/01/$ - see front matter Ó 2001 Elsevier Science B.V. All rights reserved. PII: S 0 9 2 7 - 0 2 5 6 ( 0 0 ) 0 0 1 2 3 - 3

38

J. Zhu et al. / Computational Materials Science 20 (2001) 37±47

The purpose of this paper is to develop a simple and e€ective numerical method to simulate the temporal di€usive transport through microstructures with arbitrary complexity. A microstructure is described by ®eld variables and generated by the phase-®eld method. A semi-implicit Fourier± Chebyshev spectral method was proposed to eciently solve the di€usion equation with a variable di€usion coecient. It will be shown that, using this method, both the temporal di€usion pro®le through a microstructure and the macroscopic e€ective di€usivity can be obtained. It is used to simulate the di€usion through a single-phase grain structure and two-phase microstructures with different di€usive properties in each phase. 2. Computer modeling 2.1. Microstructure description To model the di€usion through an arbitrary microstructure, we used a set of space- and timedependent ®eld variables to describe the microstructure, similar to the phase-®eld approach [4]. For example, the concentration di€erence throughout the microstructure is described by a concentration ®eld C…r†. A single-phase polycrystalline microstructure can be described by a set of order-parameter ®eld variables, g1 …r†, g2 …r†,. . ., gp …r† where p is the number of grain orientations in the system [5]. To simplify the problem, we studied di€usion in a static microstructure which remained unchanged during the di€usion process. The solid state reactions between the di€usant atoms and the microstructure such as segregation, which may occur in some practical applications, are not considered in this work. Therefore, only temporal and spatial evolution of one variable, the concentration of di€using species, is needed to investigate the diffusive transport in a microstructure described by a set of static ®eld variables. The spatial dependence of di€usivity is introduced through its dependence on the ®eld variables C…r† or g…r†, i.e., we describe the di€usion coecient D as a function of C or g written as D ˆ f …C; g1 ; g2 ; . . . ; gp †. For example, in a single-phase polycrystalline material, the di€u-

sion coecient throughout the system can P be expressed in a scaled form as D ˆ D …1 ÿ a piˆ1 g2i †,  and a are positive constants. The values where P D 2 of gi in the grain bulk are higher than those in the grain boundaries, and therefore the di€erence in grain boundary and bulk di€usion coecients can be easily described by this simple equation. Although the choice of this equation was quite arbitrary, we believed it would not a€ect the general results of the di€usive transport process. One important advantage is the fact that it avoids the speci®cation of boundary conditions at the interfaces between di€erent regions with di€erent diffusivites. Since there was no microstructural P 2 evolution in our study, the variable gi was only a function of spatial coordinates r, and thus the di€usion coecient was mapped to each lattice point as D…r†. It should also be pointed out that the ®eld variables are not required to be di€use as they are in the phase-®eld model. 2.2. Numerical method Neglecting the segregation of di€using species, the temporal di€usive transport process through the microstructure can be described by Fick's second law written as oC…r; t† ˆ r  D…r†rC…r; t†; ot

…1†

where C is the concentration of di€usant, t the time, r the spatial coordinate, and D…r† is the diffusion coecient distribution in a given microstructure. Our numerical simulation was performed in two dimensions with a periodic boundary condition in one-dimension and a ®xed boundary condition in the other which has a ®xed di€usant source and sink. The initial concentration of di€usant throughout the system is zero. Analytical solutions to the above equation are only possible with highly idealized assumptions about the di€usion coecient distribution D…r†. For studying a distribution with arbitrary complexity, numerical simulations have to be employed. Most of the existing numerical simulations for this problem were performed using explicit forward Euler method in time and ®nite-di€erence in space. To maintain the stability of the scheme

J. Zhu et al. / Computational Materials Science 20 (2001) 37±47

and to achieve high accuracy, the time step and spatial grid size have to be very small, which seriously limit the system size and time duration of a simulation. Spectral methods, which were widely used in the ®eld of ¯uid dynamics [6], o€ered us an accurate and ecient alternative to solve partial di€erential equations. Recently, Chen and Shen [7] used a semi-implicit Fourier spectral method for solving the time-dependent Ginzburg±Landau and Cahn±Hilliard equations. They showed that the semi-implicit Fourier spectral method is signi®cantly more ecient and accurate than a conventional explicit ®nite di€erence method. Consider the forward Euler method for Eq. (1) C n‡1 ÿ C n ˆ r  DrC n : Dt

…2†

The above scheme with a ®nite di€erence or Fourier approximation in space has a severe time step constraint of the form Dt 6 k=N 2 [6], where k is a constant and N is the number of lattice points to be used in each space dimension. The constraint becomes Dt 6 k=N 4 [6] when the problem has a ®xed boundary and a spectral approximation is used in space. To avoid such a severe restriction, we consider a semi-implicit treatment. The idea is to split the variable di€usion coecient D…r† into A and D…r† ÿ A, where A is a suitable constant, and treat them separately to obtain C n‡1 ÿ C n ÿ Ar2 C n‡1 ˆ ÿr  …A ÿ D†rC n : Dt

…3†

write u ˆ C n‡1 , a ˆ 1=…ADt†, and f ˆ C n =…ADt†ÿ r  …1 ÿ …D=A††rC n , and then Eq. (3) can be simpli®ed as au ÿ r2 u ˆ f ;

…4†

with a periodic boundary condition in one direction and a ®xed boundary condition in the other. At each time step, this equation can be solved accurately and eciently by a Fourier±Chebyshev Galerkin algorithm which we brie¯y describe below. Expanding u and f in Fourier series in the x direction in which a periodic boundary P condition m is applied, and substituting u…x; y† ˆ 1 mˆÿ1 u …y† eimx (same for f) in Eq. (4), we obtain

39

…a ‡ m2 †um ÿ umyy ˆ f m ; …5†

m ˆ 0; 1; 2; . . . ; um …1† ˆ 0:

Then, we solve the above equation for each m by using the Chebyshev±Galerkin algorithm developed by Shen [8]. More precisely, let Tk …y† be the kth degree Chebyshev polynomial, we look for an approximation of um in the form umN ˆ P N ÿ2 m ^k /k …y† where /k …y† ˆ Tk …y† ÿ Tk‡2 …y† are kˆ0 u the Galerkin basis functionsR satisfying /k …1† ˆ 0. 1 Let us denote …u; v†x ˆ ÿ1 uvx dy where x ˆ 2 ÿ…1=2† …1 ÿ y † is the Chebyshev weight function. N ÿ2 Then, the coecients f^ umk gkˆ0 (hence the approxm imation uN ) are determined by the weighted variational formulation for Eq. (5) …a ‡ m2 †…umN ; /k †x ÿ ……umN †yy ; /k †x ˆ …IN f m ; /k †x ;

k ˆ 0; 1; 2; . . . ; N ÿ 2;

…6†

where IN is a polynomial interpolation operator based on the Chebyshev±Gauss±Lobatto (CGL) points: yj ˆ cos…jp=N †; j ˆ 0; 1; . . . ; N . More precisely, for any continuous function g on ‰ÿ1; 1Š, IN g is the unique polynomial of degree less or equal than N such that IN g…yj † ˆ g…yj †;

j ˆ 0; 1; . . . ; N :

Furthermore, the transform between the function values ff …yj †gNjˆ0 and the Chebyshev coecients of IN f can be evaluated in O…N log N † operations by using the fast Fourier transform (FFT). Now, we describe how to determine the coeN ÿ2 cients f^ umk gkˆ0 from Eq. (6). Let us denote Z 1 T IN f /k x dy; f ˆ …f0 ; f1 ; . . . ; fN ÿ2 † ; fk ˆ ÿ1 Z 1 /00j /k x dy; S ˆ …skj †k;jˆ0;1;...;N ÿ2 ; skj ˆ ÿ ÿ1 Z 1 /j /k x dy; B ˆ …bkj †k;jˆ0;1;...;N ÿ2 : bkj ˆ ÿ1

Then, it is easy to see that Eq. (6) is equivalent to the following linear system u ˆ f; ‰…a ‡ m2 †B ‡ SŠ

…7†

40

J. Zhu et al. / Computational Materials Science 20 (2001) 37±47 T

where u ˆ …^ um0 ; u^m1 ; . . . ; u^mN ÿ2 † . It is shown in [8] that the entries of the matrices S and B are given by:

3. Results and discussion

skk ˆ 2p…k ‡ 1†…k ‡ 2†;

By numerically solving the di€usion equation with a variable coecient, we can obtain the temporal evolution of the concentration pro®le through any arbitrary microstructure. The ®rst example is the grain boundary di€usion, which has been studied for many years since the early 50s. For an idealized grain boundary geometry, analytical solutions are available for the concentration pro®les [9±11]. These solutions can be employed to analyze the grain boundary di€usion coecients from experimental measurements [12]. In recent years, some numerical models were developed to overcome the simple geometry and topology restrictions. Gui et al. [13] analyzed the grain boundary di€usion process in thin ®lms by transmission-line-matrix modeling . More recently, Swiler et al. [14,15] investigated some important heterogeneous di€usion e€ects in polycrystalline microstructures obtained from Potts model simulations. We ®rst tested our approach by studying the di€usion through an ideal grain boundary described by the Fisher's model where the analytical results were available [9]. The grain boundary is treated as a high-di€usivity, semi-in®nite isotropic slab of uniform width, embedded in a low-di€usivity, semi-in®nite perfect crystal which is normal to the surface that carries the di€usant shown in Fig. 1(a). However, in numerical simulations we can approximate a grain boundary as a ®nite slab embedded in a ®nite crystal. This approximation can represent a semi-in®nite system very well if the grain boundary thickness d is far less than the system size L (d  L). A simple way to describe such a grain boundary in Fisher's model is to choose two ®eld variables g1 , g2 . In grain 1, g1 ˆ 1, g2 ˆ 0. In grain 2 g1 ˆ 0, g2 ˆ 1. The phase ®eld model applied to describe the microstructure usually involves a di€usive interface, where the ®eld variables g1 , g2 change continuously from 0 to 1 across the grain boundary, shown in Fig. 1(b). The solid line in the Fig. 1(b) plotted the di€usion coecient D across the grain boundary described by D ˆ D ‰1 ÿ a…g21 ‡ g22 †Š where D ˆ 1

k ˆ 0; 1; . . . ; N ÿ 2;

skj ˆ 4p…k ‡ 1†; j ˆ k ‡ 2; k ‡ 4; k ‡ 6; . . . ; skj ˆ 0; j < k or j ‡ k odd; and bkk ˆ p;

k 6ˆ 0; and p bkj ˆ ÿ ; j ˆ k  2; 2 bkj ˆ 0; otherwise:

b00 ˆ 2p;

Although S is not sparse but it is a special upper-triangular matrix whose non-zero elements (except the diagonal element) at each row are all the same. Therefore, one can design a special Gaussian elimination process which takes into account this special structure to solve Eq. (7) in O…N † operations (more precisely, at a cost comparable to that of solving a penda-diagonal matrix). We refer to [8] for more details on this matter. Thus, the overall operation counts for solving Eq. (5) is O…N log N † for each m. Hence, assuming M Fourier coecients are retained in the Fourier expansion, the total cost of the Fourier± Chebyshev Galerkin method for solving Eq. (4) is O…NM log…NM†† which is quasi optimal with respect to the number of points (lattices) used. Note that the total cost is only slightly more than O…NM† operations which are required by the conventional ®nite di€erence/®nite element methods, but the Fourier±Chebyshev Galerkin method is capable of producing much more accurate results using signi®cantly less number of points (lattices). On the other hand, the semiimplicit treatment will allow us to use much larger time steps than that is allowed by an explicit method (about two orders of magnitude larger from our numerical experiments). Hence, the resulting semi-implicit Fourier±Chebyshev Galerkin method is an extremely ecient method for solving di€usion equation with a variable di€usion coecient.

3.1. Grain boundary di€usion

J. Zhu et al. / Computational Materials Science 20 (2001) 37±47

41

The kinetics of grain boundary di€usion are usually classi®ed to type A, B, and C according to the parameters of the model, such as the ratio of grain boundary di€usion coecient to the bulk lattice di€usion coecient, the di€usion time, and the ratio of grain boundary width to the grain size [16]. We have carried out simulations in all three regimes and compared the concentration pro®le at di€erent times obtained from our simulations with the analytical solutions in each regime. The A regime is for long di€usion time t and small grain size d with the condition: …Dv t†1=2  d. The concentration pro®les in type A kinetics appear to obey Fick's law for a homogeneous system with an e€ective coecient Deff which can be written as   y hCi ˆ erfc p ; …8† 2 Deff t

Fig. 1. Di€usion through an idealized grain boundary: (a) Fisher's model for grain boundary di€usion; (b) ®eld variables g1 , g2 and di€usion coecient D across a grain boundary in a phase-®eld description; (c) an isoconcentration contour plot with eight levels through a grain boundary after a certain diffusion time, D ˆ D ‰1 ÿ a…g21 ‡ g22 †Š, D ˆ 1 and a ˆ 0:999.

and a ˆ 0:999. Fig. 1(c) shows an isoconcentration contour plot through a grain boundary after a certain di€usion time. The di€usion coecient within the grain boundary being higher than in the crystal, the di€usion penetrated deeper along the grain boundary than in the grain bulk. Consequently, it started leaking through the two walls of the grain boundary into the grain bulk.

where Deff ˆ fDgb ‡ …1 ÿ f †Dv and f is the volume fraction of grain boundaries. The comparison of the average concentration hCi at di€erent times in A regime was shown in Fig. 2(a) where our numerical results (shown in symbols) agree very well with analytical solutions from Eq. (8) (shown in lines). Type C kinetics is the opposite of type A kinetics where volume di€usion in the grain bulk is 1=2 negligible, i.e., …Dv t†  d. Therefore the di€usion is only one-dimensional in Fig. 1(a) (along y direction). With the de®ned boundary condition, the concentration along the grain boundary in Fig. 1(a) is ! y …9† C…y; t† ˆ erfc p : 2 Dgb t Fig. 2(b) showed a good agreement of the numerically calculated and analytical results of the concentration along the grain boundary in regime C. Type B kinetics is of the intermediate case between B regime and C regime with the condition d  …Dv t†1=2 < d=2. Most experimental conditions fall into this regime. The analytical solution in the B regime was obtained by Whipple in a complex form [10]. The distribution of concentration through the microstructure is described by

42

J. Zhu et al. / Computational Materials Science 20 (2001) 37±47

Fig. 2. Comparison of the concentration pro®le at di€erent di€usion distance y for a grain boundary di€usion at di€erent di€usion times: (a) type A kinetics, hCi vs y; (b) type C kinetics, C…y; t† vs y; (c) type B kinetics, hCi vs y 6=5 .

  Z D g 1 g2 ‡ p exp ÿ 2 4r 2 p 1 r3=2 " r # 1 Dÿ1 rÿ1 n‡ dr:  erfc 2 Dÿr b

C ˆ erfc

g

…10†

Several dimensionless parameters pare de®ned in the above equation as: g ˆ y= Dv t , D ˆ p Dgb =Dv , p  b ˆ …D ÿ 1†d=…2 Dv t†, and n ˆ …x ÿ d=2†= Dv t. Eq. (10) was evaluated with all parameters the same as those chosen in our simulations. For example, the thickness of the grain boundary d should be determined from the di€use-interface ®eld variable pro®le [7]. Fig. 2(c) shows one example of the average concentration hCi as a function of y 6=5 at 3 di€erent time steps. We observed a good agreement between the simulation results and the analytical Whipple's solution, especially at long times in a type-B kinetics where the Whipple's solution is valid [17]. However, at short times …t ˆ 2000† there is a little di€erence between the simulated hCi and the calculated hCi because 1=2 the condition d  …Dv t† is not well satis®ed at short times. The plot of the logarithm of the average concentration versus y 6=5 was the basis for experimental measurements of grain boundary di€usion coecients [12]. Based on the simple examples discussed above, the numerical model works quite well in all three regimes of grain boundary di€usion. Although for the idealized Fisher's model, analytical solutions are possible, it is usually not the case for polycrystals, particularly if one is interested in the spatial distribution of di€using species in the microstructure. The numerical model proposed in this work can be applied to di€usion through a polycrystal without any further complication. Fig. 3 shows an example of the temporal evolution of concentration during di€usion through a polycrystalline grain structure. The grain microstructure, which was generated from a phase ®eld simulation of grain growth [5], was mapped to the x-y plane. Heterogeneous di€usion e€ect is clear that the concentration values in the grain boundaries are much higher than those in the grain bulk. As can be seen from the concentration distribution in Fig. 3(a) (t ˆ 50,000) and Fig. 3(b) (t ˆ 500,000), more species di€used into the microstructure at longer di€usion times. The di€usion coecient in the P grain microstructure is written as D ˆ D …1 ÿ a i g2i † where D and a are positive constants. Di€erent ratio of grain boundary di€usion coecient Dgb and bulk di€usion coecient Dv can be easily obtained by

J. Zhu et al. / Computational Materials Science 20 (2001) 37±47

43

Fig. 3. Temporal evolution of concentration through a polycrystalline microstructure at two di€erent times, D ˆ P D ‰1 ÿ a i g2i Š, D ˆ 1 and a ˆ 0:9999: (a) t ˆ 50,000; (b) t ˆ 500,000.

Fig. 4. Concentration pro®le in a polycrystal for two ratios ofPDgb =Dv at the same di€usion time t ˆ 10,000, D ˆ D …1ÿ a i g2i †, D ˆ 1: (a) a ˆ 0:8; (b) a ˆ 0:98.

adjusting the value of a. Fig. 4 shows the concentration pro®le through a polycrystalline material for two ratios of Dgb =Dv at the same di€usion time. The ratio in Fig. 4(b) (Dgb =Dv  20) is higher than in Fig. 4(a) (Dgb =Dv  2:5), and thus the concentration of the di€using atoms in the grain boundary di€ers more considerably from that in the grain bulk.

conductivity and di€usion coecient, were often approximated using the so-called mixture rules [18,19] X n ˆ Vi Kin ; ÿ1 6 n 6 1; …11† Keff

3.2. E€ective di€usion coecients of microstructures The e€ective properties of a microstructure, such as electrical conductivity, dielectric constant, magnetic permeability, elastic moduli, thermal

i

where Keff is the e€ective property of the microstructure, Ki and Vi are the corresponding property and volume fraction of the ith phase. More rigorous theories can predict the upper and lower bounds for the properties [2,18]. Recently, there has been increasing interest in directly calculating e€ective properties of digitized experimental microstructures or microstructures generated from

44

J. Zhu et al. / Computational Materials Science 20 (2001) 37±47

computer models [3]. Most of these numerical approaches aimed at approximate solutions of the steady-state equations using ®nite element or ®nite di€erent methods [3]. It is easy to show that one can derive the effective properties of a microstructure from the temporal concentration pro®les by solving the non-steady-state di€usion equation. We employed two approaches to extract the e€ective di€usion coecient Deff for any given microstructure. One is studying the average concentration pro®le hCi as a function of di€usion distance y at short di€usion times. Assuming Eq. (8) describes the relationship between hCi and y, we can approximately calculate Deff by ®tting di€usion data to Eq. (8). Another approach is to derive Deff from the steady-state solution of the non-steady-state di€usion equation at longer intervals of time. The ¯ux J for long di€usion time is related to the concentration gradient by Fick's ®rst law, J ˆ ÿDeff rC:

…12†

The total ¯ux J can be obtained easily from the di€usion coecient distribution D…r† and the concentration distribution C…r† through the microstructure at the steady state. rC in Eq. (12) is determined by the applied concentration gradient. Our numerical work shows Deff computed from Eqs. (8) and (12) has a di€erence less than 2%, which can satisfy most of the practical applications. One of the main advantages of this approach is that it produces both the time-dependent

concentration pro®les in a microstructure and the e€ective properties of the microstructure. In addition, it avoids any complicated boundary conditions and thus allows one to study the di€usion transport through any arbitrary microstructures. As test examples, we ®rst computed Deff for two of the simplest possible microstructures. The system is composed of parallel slabs of A and B with di€erent di€usion coecients DA and DB , shown in Fig. 5. In Fig. 5(a), the ¯ux is parallel to the plane of the slabs, and the arrangement is equivalent to a parallel electrical circuit. Each slab has the same concentration gradient, and most of the ¯ux is through the component with higher di€usivity. Deff is given by Deff ˆ DA VA ‡ DB VB , where VA and VB are the volume fraction of each component. This corresponds to n ˆ 1 in Eq. (11). In contrast, for di€usion ¯ux perpendicular to the plane of the slabs in Fig. 5(b), it is equivalent to a series electrical circuits. The ¯ux through each slab is equal, but the concentration gradients are di€erent. In this case, n ˆ ÿ1 in Eq. (11) and Deff ˆ DA DB = …DA VB ‡ DB VA †. Fig. 6 shows the comparison of calculated and theoretical Deff as a function of volume fraction for these two parallel slab microstructures. Slab arrangements in Fig. 5(a) and (b) are indicated by jj and ˆ respectively. As can be seen from the graph, the e€ective di€usion coecients computed agree well with those of theoretical solutions for both cases, thus validating our methods. Generally Deff for the two parallel slab microstructures will cover the complete range of

Fig. 5. Idealized two-phase microstructures ± parallel slabs: (a) ¯ux parallel to the plane of slabs …jj†; (b) ¯ux perpendicular to the plane of slabs ( ˆ ).

J. Zhu et al. / Computational Materials Science 20 (2001) 37±47

Fig. 6. Computed and theoretical e€ective di€usion coecient Deff as a function of volume fraction of phase A for parallel slab microstructures, DA ˆ 1.0, DB ˆ 0.2.

e€ective properties between the upper and lower bounds as shown in the solid line and dashed line in Fig. 6. It would be expected that for arbitrary microstructures, Deff was within this range. For a polycrystalline material, di€usion was carried out as described in the previous section. In particular, we examined the e€ect of the average grain size d of a grain structure on the e€ective di€usion coecient. The results are plotted as the ®lled circles shown in Fig. 7, where Deff;simulation is calculated from steady state concentration pro®le. Deff;simulation varies inversely with the average grain

Fig. 7. E€ective di€usion coecient Deff as a function of average grain size d for a single-phase polycrystalline microP structure, D ˆ 1 ÿ 0:999 i g2 . Deff;upper and Deff;low are the e€ective di€usion coecients predicted by mixture rules.

45

size d. This result is in good agreement with that of sharp-interface simulations by Swiler et al. [14]. The driving force for grain growth is the reduction of grain boundary energy. As the grain size becomes larger, the volume fraction of grain boundaries decreases, resulting in a decrease in the e€ective di€usion coecient. In Fig. 7, we also plotted the theoretical e€ective di€usion coecients Deff;upper and Deff;low as the upper and low bounds of the e€ective di€usion coecients which can be predicted by rule of mixing as Dgb fgb ‡ Dv …1 ÿ fgb † and Dgb Dv =‰Dgb …1 ÿ fgb † ‡ Dv fgb Š, respectively, where fgb is the volume fraction of grain boundaries. However, for a di€use-interface description of the microstructure, there is some ambiguity in determining the exact values of Dgb , Dv and fgb . Recognizing that Deff predicted by rule of mixing is essentially a P volume average, we calculated Deff;upper as …1=N † r D…r† where N is the total number of latticePpoints. Similarly Deff;low was computed as N =… 1=D…r††. As expected, for all the grain sizes studied, Deff;simulation is between Deff;upper and Deff;low , which are the e€ective di€usion coecients for idealized grain structures where grain boundaries can be treated as parallel slabs embedded in the grains. The di€erence between Deff;simulation and Deff;upper was explained by Swiler et al. [15] as a result of triple junctions between grains as bottlenecks to di€usion. Their sharp-interface simulations showed that Deff;simulation  0:27Dgb d=d [14], where d is the grain boundary width and d=d is the approximated volume fraction of grain boundaries. As a ®nal example, di€usion through two-phase microstructures was studied. The two-phase microstructures in Fig. 8 were obtained from our phase ®eld simulations of phase separation of an initially homogeneous system. The two-phase system forms an inter-connected or a dispersed microstructure at di€erent volume fractions. Fig. 9 shows the change of Deff (obtained from the steady state concentration pro®le) as a function of volume fraction for these microstructures (plotted with circle symbols). As the volume fraction of phase A with high di€usion coecient …DA  DB † increases, the shape of A droplets changes, and Deff increases gradually. After VA reaching about 50%, there is a sharp increase in Deff when the

46

J. Zhu et al. / Computational Materials Science 20 (2001) 37±47

Fig. 8. Two-phase microstructures for di€erent volume fraction of phase A (shown in white color) showing the percolation transition, (a)±(e) for VA ˆ 16%, 32%, 50%, 62%, and 87%, respectively.

Fig. 9. E€ective di€usion coecient Deff as a function of volume fraction. Circles were calculated from the microstructures in Fig. 8. Dashed line and solid line were plotted from Eq. (11) representing the upper and lower bounds of Deff . DA =DB ˆ 1000.

microstructure changes from isolated phase A in B to inter-connected A and B phases. The ratio of DA =DB is three orders of magnitude in Fig. 9. The dashed line and solid line represent the upper and low bounds of Deff determined from Eq. (11) as n ˆ 1 and ÿ1, respectively. It is usually observed that the larger the distinction between the properties of di€erent phases in heterogeneous systems, the more obvious is the percolation transition [18]. Our numerical experiments have demonstrated that the Fourier±Chebyshev spectral method is very ecient for solving the time-dependent diffusion equation in which the di€usivity distribution is a smooth function throughout the

microstructure, especially for microstructures described by the di€use-interface model. The excellent agreement between our di€use-interface simulation and sharp-interface analytical or numerical methods showed that di€use-interface model is very effective for describing di€usional transport in mesoscale microstructures. In recent years, the di€use-interface phase-®eld approach has been recognized as a powerful mathematical model for simulating the mesoscale morphological pattern formation such as grain growth and Ostwald ripening. By coupling the phase-®eld modeling of microstructure evolution, the approach developed in this paper allows us to model the e€ective di€usivity evolution in an evolving microstructure. More details will be reported in the near future.

4. Conclusion We have simulated the temporal mass di€usion transport through an arbitrary microstructure based on a phase-®eld approach. Di€usion coecients were calculated as a function of the order parameters describing a single-phase grain structure where grain boundary di€usion was much higher than bulk di€usion and a two-phase microstructure where the di€usivities varied with the phase. The semi-implicit Fourier±Chebyshev spectral method was shown to be accurate and ecient in solving the variable di€usion equations. The accuracy of the model was demonstrated by reproducing the classic di€usion results of grain boundary di€usion in all three kinetic regimes

J. Zhu et al. / Computational Materials Science 20 (2001) 37±47

(type A, B and C) as well as the parallel slabs geometries. We demonstrated that one can extract e€ective di€usion coecient of a microstructure from the temporal di€usion pro®les. Furthermore, the sharp interface results of Swiler et al. were reproduced to show that the e€ective di€usivity varies inversely with grain size. This model was used to predict the e€ective di€usivity of twophase percolating and non-percolating structures where each phase has very di€erent di€usivities. This approach can be also used to study similar transport problems such as heat transport by recognizing the similarity between mass di€usion and thermal di€usion equations. Acknowledgements The authors are grateful for the ®nancial support from the Sandia National Laboratory and the National Science Foundation under Grant No. DMR 96-33719 and DMS 9721413. References [1] I. Kaur, Y. Mishin, W. Gust, Fundamentals of Grain and Interphase Boundary Di€usion, third ed., Wiley, Chichester, 1995.

47

[2] S. Torquato, Int. J. Solids Struct. 35 (1998) 2385. [3] E.J. Garboczi, D.P. Bentz, Constr. Building Mater. 10 (1996) 293. [4] L.Q. Chen, Y.Z. Wang, JOM 48 (1996) 13. [5] D.N. Fan, L.Q. Chen, S.P. Chen, Mater. Sci. Eng. A 238 (1997) 78. [6] C. Canuto, M.Y. Hussaini, A. Quarteroni, T.A. Zang, Spectral Methods in Fluid Dynamics, Springer, Berlin, 1987. [7] L.Q. Chen, J. Shen, Comput. Phys. Commun. 108 (1998) 147. [8] J. Shen, SIAM J. Sci. Comput. 16 (1995) 74. [9] J.C. Fisher, J. Appl. Phys. 22 (1951) 74. [10] R.T.P. Whipple, Philos. Mag. 45 (1954) 1225. [11] T. Suzuoka, J. Phys. Soc. Jpn. 19 (1964) 839. [12] H.S. Levine, C.J. MacCallum, J. Appl. Phys. 31 (1960) 595. [13] X. Gui, S.K. Dew, M.J. Brett, J. Appl. Phys. 74 (1993) 7173. [14] T.P. Swiler, V. Tikare, E.A. Holm, Mater. Sci. Eng. A 238 (1997) 85. [15] T.P. Swiler, E.A. Holm, Ceram. Trans. 69 (1997) 203. [16] L.G. Harrison, Trans. Faraday Soc. 57 (1961) 1191. [17] J. Philebert, Atom Movements ± Di€usion and Mass Transport in Solids, Les Editions de Physique, Les Ulis, 1991. [18] C.W. Nan, Prog. Mater. Sci. 37 (1993) 1. [19] L.E. Nielsen, Predicting the Properties of Mixtures: Mixture Rules in Science and Engineering, Marcel Dekker, New York, 1978.